Synthesis and Electrochemistry of Anthraquinone

Feb 5, 1999 - Albert A. Tuinman,† David C. Baker,† and James Q. Chambers*,†. Department of Chemistry, University of Tennessee, Knoxville, Tennes...
6 downloads 0 Views 137KB Size
Bioconjugate Chem. 1999, 10, 261−270

261

Synthesis and Electrochemistry of Anthraquinone-Oligodeoxynucleotide Conjugates Neil A. Whittemore,† Adam N. Mullenix,† Gopal B. Inamati,‡ Muthiah Manoharan,‡ P. Dan Cook,‡ Albert A. Tuinman,† David C. Baker,† and James Q. Chambers*,† Department of Chemistry, University of Tennessee, Knoxville, Tennessee 37996-1600, and Chemistry Division, Isis Pharmaceuticals, Carlsbad Research Center, 2292 Faraday Avenue, Carlsbad, California 92008. Received August 19, 1998; Revised Manuscript Received December 10, 1998

Electroactive oligodeoxynucleotides (ODNs) with specific base sequences have a potential application as electrical sensors for DNA molecules. To this end, a phosphoramidite that bears a 9,10-anthraquinone (AQ) group tethered to the 2′-O of the uridine via a hexylamino linker, 2′-O-[6-[2-oxo(9,10anthraquinon-2-yl)amino]hexyl]-5′-O-(4,4′-dimethoxytrityl)uridine 3′-[2-(cyanoethyl)bis(1-methylethyl)phosphoramidite] (3), has been synthesized and used to prepare three ODNs with tethered AQs using standard phosphoramidite chemistry. The synthetic methodology thus allows the synthesis of ODNs with electroactive tags attached to given locations in the base sequence. Cyclic voltammetric behavior of these AQ-ODN conjugates was examined in aqueous buffer solutions at a hanging mercury drop electrode. At slow sweep rates, nearly reversible two-electron waves characteristic of an adsorbed anthraquinone/hydroquinone redox couple was observed for all of the AQ-ODN conjugates. Approximate Langmuirian isotherms were found for the AQ-ODNs with molecular footprints, calculated from the saturation coverages, that scaled with molecular size. The cyclic voltammetric response of the duplexes formed from the AQ-ODNs and their complementary ODN was complicated by the competitive adsorption of the individual ODNs and possibly the duplex species as well.

INTRODUCTION

The exquisite selectivity provided by Nature in the formation of DNA double helices via Watson-Crick hydrogen-bonded base pairs is arguably unmatched among molecular recognition strategies. Hybridization of complementary oligonucleotides, if confined to a small volume element or interface, is the basis of a selective nucleic acid sensor using conventional (e.g., radiochemical) means of distinguishing double-helix from singlestrand (ss) structures (1). Much of this research is highly interdisciplinary, involving both the synthesis of modified oligonucleotides and their study by different physicochemical techniques (2). The use of fluorescent oligonucleotide probes is well established and the basis of commercial DNA analyzers. Reviews of the older literature on this topic are available (3, 4), and a summary of recent attempts to improve detection of the hybridization event was given by Thiel et al. (5), who used surface plasmon resonance imaging to differentiate between ssand ds-DNA. A modified electrode surface in which an electroactive probe has been incorporated would have many attractive features common to modern electroanalytical techniques. These include in situ analysis of small volumes and possibly no requirement for washing of excess reactants from the surface reaction layer. Ideally, a surfaceattached oligodeoxynucleotide (ODN) with an attached electroactive group would mimic natural DNA, which gives well-defined faradaic signals for ss-DNA and not * To whom correspondence should be addressed. Telephone: (423) 974-3437. Fax: (423) 974-3454. E-mail: jqchambers@ utk.edu. † University of Tennessee. ‡ Isis Pharmaceuticals.

ds-DNA (6-8). An ideal tag for this purpose would give a persistent signal, preferably electrochemically reversible, in a potential region well removed from where natural DNA and RNA are electroactive, such that the signal could be easily interrogated in a sensor configuration. ODNs with attached electroactive groups or tags have been described by several groups (9-13). These applications have involved chromatographic separation of the ss- and ds-DNAs with electrochemical detection of the tag, although Xu and Bard used electrochemical luminescence to detect ds-DNA with an end-tethered Ru(bpy)32+ group (14). The above strategy can be contrasted to other electrochemical means of detecting ds-DNA that have appeared in recent years, many of which involve the use of electroactive intercalators that selectively bind to doublehelix structures. Most of these methods require the addition of a solution species, the intercalator, to perform the analysis. Thus, hybridization can be detected by the preconcentration or binding of the intercalator to a surface-confined DNA molecule, which leads to an increase in the signal. Transition metal complexes, which bind to DNA to give electroactive complexes (15-17), have been used in this manner (16, 18-21), as have a variety of other species, including a Hoechst dye (22), ethidium (23), methylene blue (24), daunomycin (25), and aromatic amines (26). Other electrochemical approaches involve the enhancement of the electrocatalytic current for oxidation of ds-DNA by intercalating redox couples (27-30), use of the Fe(CN)63-/4- couple as a permeability marker (31), and an attenuation of the poly(pyrrole) doping process by a tethered oligonucleotide in the double-helix form (32). Described below are the synthesis and electrochemical behavior of ODNs with 9,10-anthraquinone (AQ) groups

10.1021/bc980095i CCC: $18.00 © 1999 American Chemical Society Published on Web 02/05/1999

262 Bioconjugate Chem., Vol. 10, No. 2, 1999 Scheme 1

tethered to the 2′-O of a uridine unit (see Scheme 1). While this work focuses on the electrode reactions, others have prepared ODNs with tethered quinone groups for different applications. An increase in stability has been noted for nucleic acid duplexes modified at the sugar 2′position (33), including several studies in which pendant AQ groups were used (34-36). Rokita and co-workers have used tethered quinones for photochemical methylation of DNA with known sequences at specific sites (37, 38), and Gasper and Schuster found photooxidation of GG sites in a DNA double helix possessing an endtethered anthraquinone group (39). Cohen and co-workers utilized 5′-linked AQ-ODNs to demonstrate that AQ derivatives can be delivered to target nucleic acids (mRNA or DNA) (40). AQ tags were chosen for the present study since they can withstand the conditions required for the solid-phase synthesis of ODNs (33, 39), while also exhibiting a two-electron electrode reaction characteristic of the quinone/hydroquinone couple. The formal potential of this couple is well separated from potentials of redox reactions associated with the ODNs. In addition, the electrochemical signal provides a means of studying interfacial electron and proton transfer reactions of immobilized oligonucleotide probes on electrode surfaces. EXPERIMENTAL PROCEDURES

All reactions were monitored by thin-layer chromatography (TLC). Adsorption chromatography was carried out using E. Merck silica gel products: (a) TLC on 0.2 mm aluminum-backed plates or (b) column chromatography using 230-400 mesh silica gel. Visualization of the TLC plates was achieved with 254 nm UV light and by sprayheat development using a p-anisaldehydesulfuric acid reagent (41). The solvent systems for adsorption chromatography were (A) 1:1 ethyl acetate/hexanes, (B) 95:5 dichloromethane/methanol, (C) 3:1 ethyl acetate/hexanes, (D) 9:1 dicholoromethane/methanol, (E) 99:1 dichloromethane/methanol, and (F) 65:35:trace ethyl acetate/ hexanes/triethylamine. All reactions were carried out under a nitrogen atmosphere unless otherwise indicated. Solvents were evaporated under an aspirator vacuum at about 25 °C unless otherwise indicated. Elemental analyses were furnished by Atlantic Microlab, Inc. (Atlanta,

Whittemore et al.

GA). 1H and 13C NMR spectra were recorded at 400.13 and 100.61 MHz, respectively, unless stated otherwise. 19F NMR spectra were recorded on a JEOL FX90Q spectrometer at 84.24 MHz using an external standard of trifluoroacetic acid-d. 31P NMR spectra were recorded at 161.9 MHz using an external standard of phosphoric acid (85%) at 37 °C. Positive ion electrospray mass spectra of the modified nucleosides were acquired on a Quattro-II quadrupole instrument (Micromass Inc., Manchester, U.K.) by infusion (5 µL/min) of a ca. 1 µM solution of the analyte in methanol. Data were averaged in the “multichannel acquisition” mode over a 2 min acquisition period. Conversely, negative ion electrospray mass spectra of the unmodified ODNs were obtained on the same instrument by infusion (5 µL/min) of a ca. 1 µM solution of the analyte in 1:1 methanol/water that contains 0.15% triethylamine. The averaging of the data was performed as described above. The AQ-ODNs were synthesized using a Perspective Biosystems Expedite 8901 Nucleic Acid Synthesis System (PE Applied Biosystems, Foster City, CA) Preparative reversed-phase HPLC on the monotritylated AQ-ODNs was performed on a Waters 600E instrument (260 nm) equipped with a Waters Delta Pak C4 column (7.8 mm × 300 mm) and a Waters 991 integrator (Waters Corp., Milford, MA). The solvent systems used for reversed-phase HPLC were (G) 50 mM triethylammonium acetate at pH 7.0 and (H) acetonitrile. The gradient system employed was 95:5 G:H for 5 min with a linear increase to 4:6 G:H over the course of 50 min (method A) using a flow rate of 2.5 mL/min. The purity of the AQ-ODNs was determined utilizing a Waters Delta Pak C4 column (3.9 mm × 300 mm) and the same gradient system as described above at a flow rate of 1.5 mL/min (method B). Additional analysis of the AQ-ODNs was conducted with capillary gel electrophoresis (CGE). A 100 µL aliquot of the ODN (0.2 OD) was injected into a Beckman P/ACE System 5000 Capillary Gel Electrophoresis unit, coupled to a UV-vis detector (Beckman and Coulter, Inc., Fullerton, CA) that was equipped with a column (27 cm in total length with an effective length of 20 cm) of 12% linear polyacrylamide in a buffer of 7 M urea and 0.2 M bis tris borate. A voltage of 500 V/cm was supplied to the column at ca. 40 °C. A typical analysis was completed in 12 min (method A). Prior to acquisition of the negative ion electrospray mass spectra of the AQ-ODNs on a HP 59987A series II API 5989A mass spectrometer (Hewlett-Packard Co., Palo Alto, CA), ca. 0.5 OD was desalted by osmosis from water that contained ammonium hydroxide using a Millipore VS 0.025 µm filter (Millipore Corp.). After 1 h, the purified sample was then dissolved in 1:1 2-propanol/ water (100 µL). Before the sample was infused into the spectrometer, a 2.5 µL aliquot of piperidine (100 mM) was added. A Shimadzu UV-2101PC UV-vis scanning spectrophotometer was used to determine the max of the novel AQ-ODNs (Shimadzu Corp., Kyoto, Japan). Melting points of the ODNs were determined using an HP model 8452A diode array spectrophotometer equipped with an HP model 89090A Peltier temperature controller (HewlettPackard Co.). The electrochemical studies were performed using a hanging mercury drop electrode (HMDE) purchased from EG&G (Oak Ridge, TN). The electrode area was 0.0225 cm2 for most of the voltammetric experiments; details are given elsewhere (42). The reference electrode was a Ag/

Electrochemical Studies of AQ−Oligos

AgCl wire immersed directly in the test solution. A Cypress model CS-2Ra (Cypress Systems, Inc., Lawrence, KS) potentiostat was used. Calculation of the cyclic voltammograms was carried out using software supplied by CH Instruments, Inc. (Cordova, TN). Materials and Solvents. Solutions of nonmodified phosphoramidites (0.1 M) were used as purchased from PE Applied Biosystems. 5′-d(ATG CAT TCT GCC CCC AAG AG)-3′ (7) was supplied by Operon Technologies, Inc. (Alameda, CA). The other unmodified ODN 5′-d(CTC TTG GGG GCA GAA TGC AT)-3′ (6) was furnished by Isis Pharmaceuticals (Carlsbad, CA). The entire amount of the ODNs was diluted to 10 mL with sterile deionized water. Anhydrous solvents were prepared as follows. Dichloromethane and pyridine were distilled from calcium hydride at atmospheric pressure, and 1,4-dioxane was distilled from lithium aluminum hydride. 9,10-Anthraquinone-2-carboxylic Acid 2,3,4,5,6Pentafluoro Phenyl Ester (9). To a solution of 8 (2.1 g, 7.9 mmol) in dichloromethane (50 mL) at 0 °C were added DCC (0.67 g, 4.3 mmol) and 2,3,4,5,6-pentafluorophenol (0.90 g, 4.3 mmol). The suspension was allowed to warm to room temperature overnight. The crude reaction mixture was adsorbed onto silica gel and submitted to column chromatography (hexanes), affording 9 as a pale yellow solid (1.25 g, 45%): mp 197-205 °C; Rf ) 0.91 (solvent A); 1H NMR (CDCl3) δ 7.88 (m, 2H), 8.37 (m, 2H), 8.50 (d, 1H, J ) 8.15 Hz), 8.58 (dd, 1H, J ) 1.52 Hz, J ) 8.09 Hz), 9.13 (d, 1H, J ) 1.33 Hz); 19F NMR (CDCl3) δ -152.35 (d, 2F, J ) 18.31 Hz), -157.23 (t, 1F, J ) 21.98 Hz), -161.98 (dd, 2F, J ) 18.31 Hz, J ) 21.98 Hz); 13C NMR (CDCl3 and 2 drops of DMSO-d6) δ 127.19, 127.69, 129.34, 131.59, 133.13, 133.72, 134.34, 134.84, 137.01, 160.92, 181.35, 181.68. Higher-order resonances associated with the pentafluoro moiety are between δ 120 and 150. Anal. Calcd for C21H7F5O4: C, 60.30; H, 1.69. Found: C, 60.23; H, 1.71. 2′-O-[6-[2-Oxo(9,10-anthraquinon-2-yl)amino]hexyl]-5′-O-(4,4′-dimethoxytrityl)uridine (10). To a solution of 9 (1.0 g, 2.4 mmol) in dichloromethane (40 mL) at 0 °C was added pyridine (0.19 mL, 2.4 mmol). Compound 4 (43) (1.4 g, 2.2 mmol) was then added. After 50 min, the reaction mixture was allowed to warm to room temperature for a further 3 h. The crude reaction mixture was then adsorbed onto silica gel and subjected to flash chromatography (solvent C), which afforded a pale yellow, glassy material 10 (1.6 g, 83%): Rf ) 0.44 (solvent B); 1H NMR (DMSO-d ) δ 1.34 (bm, 4H), 1.54 (bm, 4H), 6 3.07-3.50 (bm, 3H, obscured by the H2O signal), 3.503.67 (overlapping resonances, 3H), 3.71 (s, 6H), 3.98 (bs, 1H), 3.91 (bs, 1H), 4.18 (t, 1H, J ) 5.23 Hz), 5.11 (bs, 1H, exchanges with D2O), 5.28 (d, 1H, J ) 8.04 Hz), 5.81 (d, 1H, J ) 3.15 Hz), 6.88 (d, 4H, J ) 8.65 Hz), 7.117.50 (overlapping resonances, 9H), 7.73 (d, 1H, J ) 8.04 Hz), 7.91 (dd, 2H, J ) 3.36 Hz, J ) 5.21 Hz), 8.04-8.41 (overlapping resonances, 4H), 8.61 (s, 1H), 8.88 (t, 1H, J ) 4.80 Hz), 11.35 (bs, 1H, exchanges with D2O); 13C NMR (DMSO-d6) δ 26.28, 28.89, 29.04, 54.99, 62.58, 68.47, 69.82, 79.12, 80.91, 82.60, 85.87, 87.14, 101.43, 113.20, 125.42, 126.75, 126.96, 127.68, 127.84, 129.73, 132.73, 132.99, 134.39, 134.56, 135.05, 135.32, 139.50, 140.14, 144.60, 150.25, 158.10, 162.95, 164.44, 182.06; ESIMS calcd for C51H48N3O11 878.3, found [M + Na]+ 902.1 and [M + K]+ 918.0. Anal. Calcd for C51H48N3O11‚2H2O: C, 66.95; H, 5.73; N, 4.59. Found: C, 66.96; H, 5.57; N, 4.44. 2′-O-[6-[2-Oxo(9,10-anthraquinon-2-yl)amino]hexyl]uridine (5). Compound 10 (0.33 g, 0.38 mmol) was stirred with a 3% (v/v) dichloroacetic acid/dichloromethane solution (7 mL) for 10 min. The crude reaction mixture

Bioconjugate Chem., Vol. 10, No. 2, 1999 263

was subjected to flash chromatography (solvent A f ethyl acetate, followed by dichloromethane f E) to afford a pale yellow amorphous powder 5 (0.13 g, 60%). Lyophilization with 1,4-dioxane afforded an off-white, fluffy solid: Rf ) 0.39 (solvent D); 1H NMR (DMSO-d6) δ 1.31 (bm, 4H), 1.53 (bm, 4H), 3.40-3.63 (overlapping resonances, 6H), 3.86 (dd, 2H, J ) 2.28 Hz, J ) 4.56 Hz), 4.06 (dd, 2H, J ) 5.28 Hz, J ) 5.9 Hz, signal collapses to a triplet upon the addition of D2O), 5.02 (d, 1H, J ) 5.9 Hz, exchanges with D2O), 5.10 (t, 1H, J ) 5.14 Hz, exchanges with D2O), 5.63 (dd, 1H, J ) 2.06 Hz, J ) 7.87 Hz), 7.76-8.10 (overlapping resonances, 3H), 8.13-8.43 (overlapping resonances, 4H), 8.64 (bd, 1H, J ) 2.54 Hz), 8.90 (t, 1H, J ) 5.08 Hz), 11.31 (s, 1H); 13C NMR (DMSOd6) δ 25.32, 26.45, 29.07, 29.23, 60.74, 68.59, 69.92, 81.34, 85.31, 86.39, 101.98, 102.05, 125.80, 125.90, 127.02, 127.37, 133.24, 133.30, 139.79, 140.65, 140.77, 163.49, 165.04, 182.46. Anal. Calcd for C30H31N3O9‚0.5C4H8O2 (1,4-dioxane): C, 61.83; H, 5.67; N, 6.76. Found: C, 61.87; H, 5.78; N, 6.60. 2′-O-[6-[2-Oxo(9,10-anthraquinon-2-yl)amino]hexyl]-5′-O-(4,4′-dimethoxytrityl)uridine 3′-[2-(Cyanoethyl)bis(1-methylethyl)phosphoramidite] (3). Compound 10 (1.12 mmol, 0.98 g, dried in vacuo for 12 h) was dissolved in dichloromethane (ca. 1 mL) that was added under a blanket of argon. To the solution were added diisopropylammonium tetrazolide (1.13 mmol, 2.0 g) and 2-(cyanoethyl)-N,N,N′,N′-tetraisopropyl phosphorodiamidite (2.25 mmol, 0.14 g). After being stirred overnight, the crude reaction mixture was purified by column chromatography (solvent F), which afforded the phosphoramidite 3 (0.750 g, 64%): Rf ) 0.36 (solvent F); 31P NMR (CDCl ) δ 150.89; FABMS calcd for C H N O P 3 60 66 5 12 1079.45, found [M + H]+ 1080.11 and [M + Na]+ 1102.23. General Procedure for the Synthesis of 2′-OModified AQ-ODNs 1a-c and 2. All of the AQ-ODNs were synthesized (2 µmol scale)sat least two batches were prepared of each modified ODNsusing standard instrument chemistry with the exception of the step of incorporating the modified uridine moiety. This was accomplished by coupling the growing oligomer twice per cycle (5 min for each cycle) utilizing an acetonitrile solution of phosphoramidite 3 (0.08-0.1 M). At the end of the synthetic sequence, the final DMT group was left intact. After treatment with ammonium hydroxide (2830%) for ca. 16 h at 55 °C, the CPG was removed by filtration through a Gelman 0.45 µm nylon acrodisc syringe filter (Pall Gelman Sciences, Ann Arbor, MI). The crude ODNs were purified employing reversed-phase HPLC (method A). Final detritylation was accomplished with acetic acid (80%) for ca. 30 min. The trityl residues and excess salt were removed by passing the AQ-ODNs through a column of Sephadex G-25 (Sigma, St. Louis, MO). Analytical details of AQ-ODNs were as follows. 5′-d(AU*G CAT TCT GCC CCC AAG AG)-3′ (1a): tR ) 21.77 min (HPLC using method B); tR ) 4.12 min (CGE using method A); ESIMS calcd for C213H262N75O121P19 m/z 6397.4, found m/z 6398.8. 5′-d(ATG CAT TCU* GCC CCC AAG AG)-3′ (1b): tR ) 19.97 min (HPLC using method B); tR ) 5.89 min (CGE using method A); ESIMS calcd for C213H262N75O121P19 m/z 6397.4, found m/z 6398.6. 5′-d(AU*G CAT TCU* GCC CCC AAG AG)-3′ (1c): tR ) 22.84 min (HPLC using method B); tR ) 8.16 min (CGE using method A); ESIMS calcd for C233H279N76O125P19 m/z 6732.7, found m/z 6732.9. 5′-d(GAU*CT)-3′ (2): tR ) 34.72 min (HPLC using method B); tR ) 7.98 min (CGE using method A); 1H NMR (D2O) δ 5.84-5.88 (overlapping resonances, 3H), 5.98 (d,

264 Bioconjugate Chem., Vol. 10, No. 2, 1999

1H, J ) 7.55 Hz), 6.13 (d, 1H, J ) 6.45 Hz), 6.10-6.30 (overlapping resonances, 3H), 7.53 (d, 1H, J ) 6.59 Hz), 7.74 (s, 1H), 7.80-7.90 (overlapping resonances, 4H), 8.07-8.15 (overlapping resonances, 3H), 8.20 (s, 1H), 8.24 (s, 1H), 8.34 (d, 1H, J ) 1.78 Hz); 31P NMR (D2O) δ -0.32, -0.28, -0.11, 0.11; ESIMS calcd for C69H80N18O33P4 m/z 1813.4, found m/z 1813.2. Determination of the Concentration of the Novel ODNs 1a-c. To accurately determine the concentration of AQ-ODNs 1a-c, a stock solution of 20-60 ODs of each of the modified oligomers (dependent on the amount available) was prepared in sterile, deionized water (10 mL). The contribution of AQ to the 260 for modified AQODN 1a was added to the 260 for 5′-AUG CAT TCT GCC CCC AAG AG-3′, which was calculated from the nearestneighbor effect (44). The 260 of AQ was determined to be 26 400 M-1 cm-1 by subtracting the 260 of uridine from the 260 of monomer 5. Thus, the 260 of 1a was determined to be 214 000 M-1 cm-1 (260 nm). By measuring the absorbance at 260 nm, we found the concentration of the stock solution of 1a to be 2.09 × 10-5 M. Similarly, the concentrations of the stock solutions of 6 (260 ) 194 000 M-1 cm-1) and 7 (260 ) 188 000 M-1 cm-1) were determined to be 2.04 × 10-5 and 2.83 × 10-5 M, respectively. Electrochemical Studies. Cyclic voltammetry (CV) was carried out using aqueous buffers containing NaCl at various concentrations. For the pH studies on compounds 2 and 5 (see below), 0.125 M phosphate and acetate buffers were used for pH 7. The latter buffer system was also used for the determination of the adsorption isotherms at pH 7.15. Before measurements were taken, solutions were purged with purified nitrogen that had been passed through a column containing Mn(II) oxide dispersed on a vermiculite support to remove oxygen. Hybridization Conditions. Hybridization solutions were prepared by mixing equimolar aliquots (ca. 50 µL) of the stock solutions with 2-3 mL of buffer solution. Temperature profiles were obtained by heating the cell to 90 °C for 10 min, followed by slow cooling in 1 °C increments while measuring the absorbance at 260 nm. Once confirmation of duplex formation was achieved, CV was performed on the solution as described above. Initial experiments used either the pH 7.15 BR buffer solution containing 0.25 M NaCl and 3.6 mM EDTA or a 0.01 M ammonium acetate/0.001 M ammonium chloride buffer. The ammonium acetate buffers were chosen to mimic the conditions of Bayer et al. (46), who detected duplexes in solution using electrospray mass spectroscopy (albeit at higher concentrations). The latter experiments were performed in 0.1 M phosphate buffer (pH ∼7) containing 0.05 M NaCl and 5 mM EDTA. In the latter procedure, melting curves were obtained for solutions before exposure to the HMDE conditions, and the voltammetric behavior was compared to that of an identical solution that was not thermally cycled. Spurious voltammetric waves were obtained for solutions of 1a and 6 that had been exposed to the HMDE conditions before thermal cycling. RESULTS AND DISCUSSION

Synthesis of 2′-O-Modified ODNs. Entry into the novel oligomers 1a-c and 2 was crucial for this study (refer to Scheme 1). The key intermediate in the synthetic sequence was the electroactive phosphoramidite 3 (Scheme 2). Additionally, the strategy provided access to the free nucleoside 5 that was required to establish the experi-

Whittemore et al. Scheme 2

Scheme 3a

a (a) DCC, 2,3,4,5,6-pentafluorophenol, CH Cl , 0 °C f room 2 2 temperature over the course of 12 h; (b) 4, pyr, CH2Cl2, 0 °C for 50 min, room temperature for 3 h; (c) 3% (v/v) Cl2CHCOOH in CH2Cl2, 10 min; (d) diisopropylammonium tetrazolide, 2-(cyanoethyl)-N,N′,N′,N-tetraisopropyl phosphoramidite, CH2Cl2, room temperature for 12 h; (e) automated ODN synthesizer.

mental conditions required to observe the distinctive electrochemical signature of the quinone moiety. Initially, the 3-(3-methyl-1,4-naphthoquinon-2-ylsulfanyl)propionyl reporter group of Chatterjee and Rokita (38) was employed. However, when the precursor to the conjugates was subjected to concentrated ammonia, to mimic the final deblocking step in the synthesis of ODNs (47), deleterious side reactions took place. Accordingly, an AQ tag was employed in this work since these conjugates are known to survive the conditions of the oligonucleotide synthesis (34, 40). Compound 9 was synthesized utilizing DCC to condense 8 with 2,3,4,5,6-pentafluorophenol (Scheme 3). Addition of 9 to 4 (43) in the presence of pyridine secured the modified nucleoside 10. With this essential intermediate in hand, (a) phosphoramidite 3 was prepared using

Electrochemical Studies of AQ−Oligos

Bioconjugate Chem., Vol. 10, No. 2, 1999 265

Figure 2. Surface coverage vs exposure time of the HMDE drop to 0.4 µM 1a BR buffer (pH 7.15).

Figure 1. Typical cyclic voltammograms (O) of compounds (A) 5 (pH 7.15), (B) 2 (pH 7.77), and (C) 1a (pH 7.15) at 100 mV/s with accompanying EE mechanism simulations (solid lines). Simulation parameters were as follows: (A) ks1 ) 100 s-1, ks2 ) 20 s-1, E°1 ) -0.475 V, and E°2 ) -0.510 V; (B) ks1 ) 100 s-1, ks2 ) 2 s-1, E°1 ) -0.510 V, and E°2 ) -0.490 V; and (C) ks1 ) 100 s-1, ks2 ) 5 s-1, E°1 ) -0.470 V, and E°2 ) -0.470 V.

standard methodology (43) and (b) removal of the 4,4′dimethoxyltrityl (DMT) group was readily achieved with a 3% (v/v) dichloroacetic acid in CH2Cl2 (47) solution to afford free ribonucleoside 5. A standard protocol was performed to obtain AQ-ODNs 1a-c and 2 from 3 on an automated oligonucleotide synthesizer (48). Prior to removal of the final DMT group, the resulting oligomers were purified using reversed-phase HPLC. Cyclic Voltammetry. Before studies were carried out on oligomers 1a-c, cyclic voltammetry was performed on aqueous buffer solutions of 2 and 5. Immediately upon extrusion of a HMDE into a dilute aqueous solution of these AQ-ODNs, the presence of an adsorbed layer of electroactive AQ was indicated by the appearance of a voltammetric wave with a wave shape indicative of a redox process for a surface-confined species. Typical voltammograms are shown in Figure 1. These voltammograms are similar to those reported by Forster, in both position and shape, for related simple AQs on HMDEs (49). Thus, it appears that the AQ groups readily adsorb on and remain in close electrical contact with the mercury surface despite the presence of the oligonucleotide substitutent. The cyclic voltammograms of 20-mers 1a-c were similar to those of 2 and 5, except that the magnitudes of the current and the corresponding surface charge were

smaller. Figure 1C shows a typical cyclic voltammogram for 1a, which has an AQ group tethered to the second nucleotide from the 5′-phosphate end of the base sequence. The shapes of the voltammetric waves for all the compounds studied at sweep rates of less than ca. 1 V/s were symmetrical, or nearly so, with peak separations (Epanod - Epcath) of a few millivolts and minimal faradaic current at potentials on the wings of the voltammogram. These observations are consistent with formation of a self-assembled layer on the electrode surface. It will be argued below that these films must possess a fair degree of order as well. The voltammograms in Figure 1, which do not have ideal Nernstian shapes for an n-electron process, have been fit to a simple global EE mechanism (i.e., two successive electron additions) (50). This approach gave good fits to the slow sweep rate voltammograms of all the conjugates, including the splitting that was evident in the cyclic voltammograms at pH values of less than ca. 6 and the narrowing of the voltammograms at pH >7. In this simple analysis, it is assumed that the electron additions would be accompanied by protonation steps. It is realized that the formal potentials and rate constants used in these calculations are necessarily not unique. A more detailed analysis of the electrode kinetics for these surface layers will be published elsewhere. With time after the extrusion of a new mercury surface, the magnitude of the waves for all the AQ species increased, and then leveled off, reaching a plateau after 5-10 min, while maintaining an approximately constant shape. This behavior is shown in Figure 2, where the surface coverage was calculated using the charge necessary to reduce the adsorbed species from Faraday’s law in the usual fashion. At the sweep rates of these experiments, an n value of 2 was assumed for the reduction of the AQ moiety.

AQ + 2H+ + 2e- a AQH2

(1)

The solutions used to obtain the data of Figure 2 contained an excess of EDTA to scavenge metal ions, in particular adventitious Hg(II) species which are difficult to eliminate in HMDE voltammetry. The cathodic current evident as a shoulder on the negative side of the peak voltammogram of 1c in Figure 1 is likely due to the reduction of accumulated Hg(II). In the absence of EDTA, the cathodic charge continued to increase with time, and the wave shape changed, indicative of overlapping redox processes. This behavior is attributed to the reduction of the Hg(II) species accumulated by the oligonucleotide

266 Bioconjugate Chem., Vol. 10, No. 2, 1999

Whittemore et al.

Figure 4. Adsorption isotherm of compound 5 in pH 7.15 BR buffer with 0.25 M NaCl (O). The line is a calculated Langmuir isotherm where Γmax ) 0.369 nmol cm-2 and β ) 1.01 × 105 M-1. Figure 3. Average peak potential [(Epka + Epkc)/2] pH study for compound 5.

surface layer. The ability of nucleic acids to coordinate mercury species is well documented (51). The pH dependence of the E°′ [)(Epka + Epkc)/2] for the AQ/AQH2 couple of compound 5 is shown in Figure 3. The response is close to 60 mV/pH unit over a wide pH range. This is consistent with the two-electron, twoproton stoichiometry of eq 1 and with the Pourbaix diagrams of simple anthraquinones (52). A similar behavior was obtained for pentamer 2. Due to the limited amounts of the 20-mers that are available, full pH studies were not carried out for these compounds, but the peak potential data were consistent with Figure 3 and eq 1. The fact that the apparent E°′ values exhibited a lack of gross dispersion indicates that the surface layers possess a fair degree of order with respect to the location of the AQ/AQH2 couple. Adsorption Isotherms. The concentration dependence of the voltammograms for all of the AQ-ODN conjugates was examined in neutral pH phosphate buffer solutions. Limiting saturation surface coverages were seen at concentrations in the micromolar range, indicating a high affinity of all of the compounds for the mercury electrode surface. This is not surprising in view of the known affinity of the purine and pyrimidine bases, native DNA and RNA (7), and the AQ unit itself for mercury electrode surfaces. Figures 4 and 5 show the Γ versus concentration behavior for compounds 5 and 1a, respectively, along with fits to the simple Langmuir adsorption isotherm model. Although the agreement with the Langmuir model is not perfect, the fits are good enough to provide a means of comparing the AQ-ODN conjugate voltammetry as a function of chain length and degree of substitution. First, the molecular footprints calculated from the limiting Γmax values

footprint ) (ΓmaxNav x 10-14)-1 nm2/molecule (2) scale with the size of the n-mers and the degree of substitution; see Table 1. The sizes of the footprints per substituted uridine ring indicate a closely packed surface layer. The footprint calculation of eq 2 assumes that the n value of the electrode reaction is two for compounds 1a, 1b, 2, and 5, and four for compound 1c, which possesses

Figure 5. Adsorption isotherm of compound 1a in pH 7.15 BR buffer with 0.25 M NaCl (O). The line is a calculated Langmuir isotherm where Γmax ) 15.7 pmol cm-2 and β ) 1.7 × 107 M-1. Table 1. Langmuir Adsorption Isotherm Parameters for AQ Conjugates at Hg in pH 7.1 Buffer Solution

2 5 1a 1b 1c

footprint (nm2/molecule)

β (M-1)

∆Gads (kJ/mol)

0.48 1.99 11.0 10.5 11.1

1.01 × 105 7.9 × 107 8.50 × 106 2.69 × 107 3.2 × 107

-28 -45 -40 -42 -43

two pendant AQ groups. Significantly, the footprint for the latter was in agreement with those of the other 20mers, indicating that both AQ groups were involved in the electrochemical reactions. At faster sweep rates, a splitting of the cathodic wave was seen in the voltammograms of the disubstituted 20-mer. The packing density of 5, ca. 0.5 nm2/molecule, is on the order of magnitude expected for a much smaller benzohydroquinone based on van der Waals radii (53). This indicates that the AQ group in the adsorbate is oriented in an edge-pendant configuration with respect to the electrode surface. The observed reversibility of the electrode reaction also suggests that the AQ groups are close to the mercury surface. It should be noted that the charge under the voltammetric waves decreased significantly at faster sweep rates (up to 30-50% of the charge for the voltammograms of Figure 1). This is indicative of kinetic limitations on the electrode reaction which are beyond the scope of this report. However, possible kinetic limitations would decrease the surface charge and in-

Electrochemical Studies of AQ−Oligos

Bioconjugate Chem., Vol. 10, No. 2, 1999 267

Figure 6. Cyclic voltammogram of (b) 1a and (O) 1a and complementary strand 6 in 0.01 M NH4Ac and 0.001 M NH4Cl, with a sweep rate of 0.1 V s-1, a 0.0225 cm2 HMDE, and an accumulation time of 5 min.

Figure 7. Competitive adsorption study of compounds 1a and 7 at a HMDE. Conditions were 0.216 µM 1a in pH 7.15 BR buffer, with 0.25 M NaCl and 3.6 mM EDTA: (O) cathodic charge or (3) anodic charge.

crease the calculated values of the footprints, which assume a two-electron n value. The Langmuirian behavior of the Γ data also affords a quantitative measure of the affinity of the molecules for the electrode surface via the following relationship between the free energy of adsorption (∆Gads) and the Langmuir energy coefficient (β):

∆Gads ) -RT ln(β)

(3)

The large absolute values of the negative free energy of adsorption for the AQ bioconjugates indicate a strong affinity of the mercury surface for these single-strand DNA molecules. Competitive Adsorption of ODNs. Several experiments were performed to examine the effect of hybridization of the DNA chain on the voltammetric response. The results from one such experiment are shown in Figure 6, which shows the voltammogram of 1a in the absence and presence of an equimolar concentration of its complementary strand, compound 6, in an ammonium acetate buffer solution. In the presence of the complementary strand, the electrochemical signal was attenuated significantly, while the wave shape remained approximately constant. However, this result is complicated by the competitive adsorption displacement of the tagged 20mer from the electrode surface by the untagged chain. In the buffer solution of Figure 6, essentially identical behavior was observed when a noncomplementary sequence, compound 7, was added to the solution. The facile competitive nature of the adsorption process is demonstrated by the experiment whose results are depicted in Figure 7. Here the voltammetric charge for 1a was measured as a function of the amount of added untagged 20-mer in phosphate buffer solution. Within a few seconds, apparent equilibrium surface coverage by both compounds is observed. The data are in agreement with the following equation for competitive Langmuirian adsorption of two species with identical saturation surface coverages (54).

Γ1 )

Γsβ1C1 1 + β1C1 + β2C2

(4)

This equation can be rearranged into the following linear form

Figure 8. Linear competitive adsorption model.

β2C2 1 1 -1) + θ1 β1C1 β1C1

(5)

where C1 is the concentration of the tagged 20-mer 1a, C2 is the concentration of the added untagged 20-mers (6 or 7), and θ1 is the fractional surface coverage of the tagged compound. The corresponding plot is shown in Figure 8. Calculation of the relative β values from the slope of the line in Figure 8 reveals that the untagged 20-mer is adsorbed less strongly than the tagged 20-mer by 4 kJ/mol. Effect of Duplex Formation. The surface state of these AQ-ODNs in the presence of its complementary strand, i.e., adsorbed ss-DNA versus adsorbed ds-DNA, is difficult to establish in the absence of additional information. The possibility exists that a duplex structure of these ODNs would unzip and adsorb on the mercury electrode in the single-strand form. There is ample precedent for this in the classical work of Palecek on native DNA electrochemistry (7), and the large negative free energy of adsorption could provide the driving force for this process. In addition, the crowded packing of these molecules on the electrode surface may not allow enough space for surface hybridization (55). To confirm that a duplex was present in the solutions containing both a tagged 20-mer and its complementary strand, the melting/annealing process was followed by monitoring the UV absorption spectrum as a function of temperature. Figure 9 shows a cooling absorbance trace for a solution of 1a and its complementary strand, which exhibits a melting point at 66 °C. Interestingly, the

268 Bioconjugate Chem., Vol. 10, No. 2, 1999

Figure 9. Typical melting curve for 0.35 µM 1a and 6 in pH 6.83 0.1 M phosphate buffer, with 0.1 M NaCl.

Figure 10. Cyclic voltammogram of 1a with complementary 20-mer 6: (A) 0.695 µM 1a in pH 7.4 phosphate buffer (0.1 M), with 0.05 M NaCl and 5 mM EDTA; (B) addition of an equal amount of compound 6 at room temperature; and (C) after heating to 90 °C and cooling to room temperature. The sweep rate was 0.1 V s-1.

melting temperatures of the untagged duplex, i.e., 6‚7, and the duplexes of the AQ-ODNs were approximately the same, indicating that the tethered AQ group is not intercalated in the duplex structure. In pH 7.4, 0.1 M phosphate buffer containing 0.05 M NaCl and 5 mM EDTA, the Tm (∆Tm) values were 66.4, 68.3 (1.9), 68.4 (2.0), and 69.1 (2.7) °C for the duplexes of 6 with 7, 1a, 1b, and 1c, respectively. Somewhat larger ∆Tm values were seen for 2′-O-modified ribonucleotides with an intercalating substituent, 2-aminoanthraquinone, tethered on shorter linker arms (35). Cyclic voltammetry experiments carried out in parallel with the melting curves gave further support to the competitive nature of the adsorption equilibria that are at play at the electrode surface. In these experiments, solutions containing equimolar amounts of a tagged AQODN and either its complementary (6) or a noncomplementary (7) ODN species were split into two parts and subjected to both CV and UV-vis analysis. The resulting voltammograms for compound 1a are shown in Figures 10 and 11 for a pH 7.4 phosphate buffer. The cyclic voltammograms marked A in these figures were obtained with solutions containing 1a at a concentration sufficient to give a limiting surface coverage of the AQ-ODN. Addition of the complementary strand to the solution resulted in an attenuation of the signal; refer to the cyclic voltammogram marked B in Figure 10. However, this result cannot be distinguished from simple competitive adsorption of an untagged ODN since essentially identi-

Whittemore et al.

Figure 11. Cyclic voltammogram of 1a with noncomplementary 20-mer 7. Conditions were the same as those described in the legend of Figure 10.

cal behavior was seen when the experiment was repeated with a noncomplementary sequence; see the cyclic voltammogram marked B in Figure 11. Significantly, however, when the solution containing 1a and its complement was annealed at 90 °C, followed by slow cooling to room temperature, the resulting cyclic voltammogram obtained on the solution was further attenuated and shifted to slightly more negative potential; see the cyclic voltammogram marked C in Figure 10. The corresponding cyclic voltammogram obtained for the control solution with the noncomplementary species 7 (cyclic voltammogram marked C in Figure 11) was not attenuated. These results are best understood if the hybridization is slow in these dilute solutions at room temperature. In this case, the attenuation seen in the cyclic voltammograms marked B in Figures 10 and 11 could be due to competitive adsorption as described above. While the further attenuation seen in the cyclic voltammogram marked C in Figure 10 appears to be related to formation of the duplex species in solution, the physical phenomenon responsible for this effect is not clear. One possibility is that the duplex 1a‚6 species is electroinactive or not adsorbed on the electrode surface and that the signal obtained after solution hybridization is due to adsorbed 1a in equilibrium with a decreased concentration of 1a in the bulk of the solution. Alternatively, the duplex 1a‚6 might adsorb on the electrode with a footprint that is larger than the combined footprints of the tagged and untagged 20-mers. More likely, the duplex 1a‚6 is electroactive, but has a smaller adsorption coefficient (β) than the single-strand ODNs since the DNA bases are not free to interact with the surface, so that when the duplex is formed in solution, the surface layer is no longer at saturation coverage. In any case, the possibility of a simple in situ electrochemical detection of the hybridization event based on these results is problematic without a different strategy and further modification of the electrode surface. CONCLUSIONS

(1) We have demonstrated that AQ-ODNs, which bear one or more reporter groups, can be prepared with a modified uridine moiety in any location in a specific sequence utilizing standard phosphoramidite methodology. (2) The AQ conjugates form compact surface layers on Hg electrodes with molecular footprints, calculated from the saturation coverages, that scale with molecular size.

Electrochemical Studies of AQ−Oligos

The cyclic voltammetry is consistent with rapid electron transfer and little dispersion of E°′ values. This indicates that the AQ groups are located in a uniform position in a compact surface layer and are not insulated from the electrode surface since facile transfer of electrons through the adsorbed ss-DNA layer is unlikely. (3) The cyclic voltammetric response of the ds-AQ-20mers is complicated by the competitive adsorption of the individual ss-20-mers and possibly the ds-20-mers as well. At room temperature in aqueous buffer solutions, equilibrium surface coverages are established more rapidly than the hybridization kinetics, which are much slower than the CV time scale. (4) Accumulation of adventitious Hg(II) ions is a problem, which has not been fully realized in previous work, when HMDEs are used for voltammetric sensing of DNA. ACKNOWLEDGMENT

This research was supported by the National Science Foundation (NSF Grant CHE-9616994) and The University of Tennessee (Knoxville, TN). LITERATURE CITED (1) Chee, M., Yang, R., Hubbell, E., Berno, A., Huang, X. C., Stern, D., Winkler, J., Lockhart, D. G., Morris, M. S., and Fodor, S. P. A. (1996) Accessing genetic information with high-density DNA arrays. Science 274, 610-614. (2) Kessler, C. (1992) Nonradioactive Labeling and Detection of Biomolecules, Springer-Verlag, Berlin. (3) Englisch, U., and Gauss, D. H. (1991) Chemically Modified Oligonucleotides. Angew. Chem., Int. Ed. 30, 613-629. (4) Asseline, U., Thoung, N. T., and Helene, C. (1997) Synthesis and properties of oligonucleotides covalently linked to intercalating agents. New J. Chem. 21, 5-17. (5) Thiel, A. J., Frutos, A. G., Jordan, C. E., Corn, R. M., and Smith, L. M. (1997) In Situ Surface Plasmon Resonance Imaging Detection of DNA Hybridization to Oligonucleotide Arrays on Gold Surfaces. Anal. Chem. 69, 4948-4956. (6) Palecek, E. (1983) Modern Polarographic (Voltammetric) Techniques in Biochemistry and Molecular Biology. In Topics in Bioelectrochemistry and Bioenergetics (G. Milazzo, Ed.) Vol. 5, pp 65-155, Wiley, New York. (7) Palecek, E. (1988) New Trends in Electrochemical Analysis of Nucleic Acids. Bioelectrochem. Bioenerg. 20, 179-194. (8) Hall, J. M., Moore-Smith, J., Bannister, J. V., and Higgins, I. J. (1994) An Electrochemical Method for Detection of Nucleic Acid Hybridisation. Biochem. Mol. Biol. Internat. 32, 21-28. (9) Jacobson, K. B., Arlinghus, H. F., Schmitt, H. W., Sachleben, R. A., Brown, G. M., Thonnard, N., Sloop, F. V., Foote, R. S., Larimer, F. W., et al. (1991) An Approach to the Use of Stable Isotopes for DNA Sequencing. Genomics 9, 51-59. (10) Sloop, F. V., Brown, G. M., Sachleben, R. A., Garrity, M. L., Elbert, J. E., and Jacobson, K. B. (1994) Metalloorganic Labels for DNA Sequencing and Mapping. New J. Chem. 18, 317-326. (11) Takenaka, S., Uto, Y., Kondo, H., Ihara, T., and Takagi, M. (1994) Electrochemically Active DNA Probes: Detection of Target DNA Sequences at Femtomole Level by HPLC with Electrochemical Detection. Anal. Biochem. 218, 436-443. (12) Ihara, T., Maruo, Y., Takenaka, S., and Takagi, M. (1996) Ferrocene-oligonucleotide conjugates for electrochemical probing of DNA. Nucleic Acids Res. 24, 4273-4281. (13) Uto, Y., Kondo, H., Abe, M., Suzuki, T., and Takenaka, S. (1997) Electrochemical analysis of DNA amplified by the polymerase chain reaction with a ferrocenylated oligonucleotide. Anal. Biochem. 250, 122-124. (14) Xu, X.-H., and Bard, A. J. (1995) Immobilization and Hybridization of DNA on an Aluminum(III) Alkanebisphosphonate Thin Film with Electrogenerated Chemiluminescent Detection. J. Am. Chem. Soc. 117, 2627.

Bioconjugate Chem., Vol. 10, No. 2, 1999 269 (15) Carter, M. T., Rodriquez, M., and Bard, A. J. (1989) Voltammetric Studies of the Interaction of Metal Chelates with DNA. 2. Tris-Chelated Complexes of Cobalt(III) and Iron(II) with 1,10-Phenanthroline and 2,2-Bipyridine. J. Am. Chem. Soc 111, 8901. (16) Rodriguez, M., and Bard, A. J. (1990) Electrochemical Studies of the Interaction of Metal Chelates with DNA. 4. Voltammetric and Electrogenerated Chemiluminescent Studies of the Interaction of Tris(2,2-bipyridine)osmium(II) with DNA. Anal. Chem. 62, 2658. (17) Welch, T. W., Corbett, A. H., and Thorp, H. H. (1995) Electrochemical Determination of Nucleic Acid Diffusion Coefficients through Noncovalent Association of Redox-Active Probes. J. Phys. Chem. 99, 11757-11763. (18) Millan, K. M., Saraullo, A., and Mikkelsen, S. R. (1994) Voltammetric DNA Biosensor for Cystic Fibrosis Based on a Modified Carbon Paste Electrode. Anal. Chem. 66, 29432948. (19) Wang, J., Rivas, G., Parrado, C., Cai, X., and Flair, M. N. (1997) Electrochemical biosensor for detecting DNA sequences from the pathogenic protozoan Cryptosporidium parvum. Talanta 44, 2003-2010. (20) Wang, J., Rivas, G., Cai, X. H., Dontha, N., Shiraishi, H., Luo, D. B., and Valera, F. S. (1997) Sequence-specific electrochemical biosensing of M. tuberculosis DNA. Anal. Chim. Acta 337, 41-48. (21) Zhao, Y.-D., Pang, D.-W., Wang, Z.-L., Cheng, J.-K., and Qi, Y.-P. (1997) DNA-modified electrodes. Part 2. Electrochemical characterization of gold electrodes modified with DNA. J. Electroanal. Chem., 203-209. (22) Hasshimoto, K., Ito, K., and Ishimori, Y. (1994) Sequencespecific gene detection with a gold electrode modified with DNA probes and an electrochemically active dye. Anal. Chem. 66, 3830. (23) Liu, S., Ye, J., He, P., and Fang, Y. (1996) Voltammetric determination of sequence-specific DNA by electroactive intercalator on graphite electrode. Anal. Chim. Acta 335, 239. (24) Kelley, S. O., Barton, J. K., Jackson, N. M., and Hill, M. G. (1997) Electrochemistry of Methylene Blue Bound to a DNA-Modified Electrode. Bioconjugate Chem. 8, 31-37. (25) Palanti, S., Marrazza, G., and Mascini, M. (1996) Electrochemical DNA probes. Anal. Lett. 29, 2309-2331. (26) Wang, J., Rivas, G., Luo, D., Cai, X., Valera, F. S., and Dontha, N. (1996) DNA-Modified Electrode for the Detection of Aromatic Amines. Anal. Chem. 68, 4365. (27) Johnston, D. H., Glasgow, K. C., and Thorp, H. H. (1995) Electrochemical Measurement of the Solvent Accessibility of Nucleobases Using Electron Transfer between DNA and Metal Complexes. J. Am. Chem. Soc. 117, 8933-8938. (28) Johnston, D. H., and Thorp, H. H. (1996) Cyclic Voltammetry Studies of Polynucleotide Binding and Oxidation by Metal Complexes: Homogeneous Electron-Transfer Kinetics. J. Phys. Chem. 100, 13837-13843. (29) Napier, M. E., and Thorp, H. H. (1997) Modification of Electrodes with Dicarboxylate Self-Assembled Monolayers for Attachment and Detection of Nucleic Acids. Langmuir 13, 6342. (30) Napier, M. E., Loomis, C. R., Sistare, M. F., Kim, J., Eckhardt, A. E., and Thorp, H. H. (1997) Probing Biomolecule Recognition with Electron Transfer: Electrochemical Sensors for DNA Hybridization. Bioconjugate Chem. 8, 906-913. (31) Tohda, K., Ameniya, S., and Umezawa, Y. (1997) Channel Mimetic Sensing Membranes for Nucleotides Based on Multitopic Hydrogen Bonding. Isr. J. Chem. 37, 267-275. (32) Korri-Youssoufi, H., Garnier, F., Srivastava, P., Dodillot, P., and Yassar, A. (1997) Toward Bioelectronics: Specific DNA Recognition Based on an Oligonucleotide-Functionalized Polypyrrole. J. Am. Chem. Soc. 119, 7388-7389. (33) Freier, S. M., and Altmann, K.-H. (1997) The ups and downs of nucleic acid duplex stability: structure-stability studies on chemically-modified DNA:RNA duplexes. Nucleic Acids Res. 25, 4429. (34) Yamana, K., Nishijima, Y., Ikeda, T., Gokota, T., Ozaki, H., Nakano, H., Sangen, O., and Shimidzu, T. (1990) Synthesis and Interactive Propeties of an Oligonucleotide with

270 Bioconjugate Chem., Vol. 10, No. 2, 1999 Anthraquinone at the Sugar Fragment. Bioconjugate Chem. 1, 319-324. (35) Keller, T. H., and Ha¨ner, R. (1993) Synthesis and hybridization properties of oligonucleotides containing 2′-O-modified ribonucleotides. Nucleic Acids Res. 21, 4499-4505. (36) Deshmukh, H. M., Joglekar, S. P., and Broom, A. D. (1995) Self-Complementary Oligodeoxyribonucleotides Containing 2′-O-(Anthraquinone-2-methyl)adenosine. Bioconjugate Chem. 6, 578-586. (37) Kang, H., and Rokita, S. E. (1996) Site-specific and photoinduced alkylation of DNA by a dimethylanthraquinoneoligodeoxynucleotide conjugate. Nucleic Acids Res. 24, 38963902. (38) Chatterjee and Rokita, S. E. (1994) The Role of a Quinone Methide in the Sequence Specific Alkylation of DNA. J. Am. Chem. Soc. 116, 1690-1697. (39) Gasper, S. M., and Schuster, G. G. (1997) Intramolecular Photoinduced Electron Transfer to Anthraquinone Linked to Duplex DNA: The Effect of Gaps and Traps on Long-Range radical Cation Migration. J. Am. Chem. Soc. 119, 1276212771. (40) Mori, K., Subasinghe, C., and Cohen, J. S. (1989) Oligodeoxynucleotide analogs with 5-linked anthraquinone. FEBS Lett. 249, 213-218. (41) Schaumberg, J. P., Hokanson, G. C., French, J. C., Smal, E., and Baker, D. C. (1985) 2′-Chloropentostatin, a new inhibitor of adenosine deaminase. J. Org. Chem 50, 1651. (42) Wagner, G. J., and Chambers, J. Q. (1997) Electrochemistry of Covalently Mercurated Uridine Nucleotides. Langmuir 13, 3529-3541. (43) Manoharan, M., Tivel, K. L., Andrade, L. K., and Cook, P. D. (1995) 2-O- and 3-O-Pyrimidine Aminotether-Containing Oligonucleotides: Synthesis and Conjugation Chemistry. Tetrahedron Lett. 36, 3647-3650. (44) Fasman, G. D. (1985) Handbook of Biochemistry and Molecular Biology: Nucleic Acids, 2nd ed., Vol. 1, p 589, CRC Press, Cleveland, OH. (45) Dean, J. A. (1985) Lange’s Handbook of Chemistry, 13th ed., McGraw-Hill, New York.

Whittemore et al. (46) Bayer, E., Bauer, T., Schmeer, K., Bleicher, K., Maier, M., and Gaus, H.-J. (1994) Analysis of Double-Stranded Oligonucleotides by Electrospray Mass Spectrometry. Anal. Chem. 66, 3858-3863. (47) Atkinson, T., and Smith, M. (1984) Solid-Phase Synthesis of Oligodeoxynucleotides by the Phosphite-triester Method. In Oligonucleotide Synthesis: A Practical Approach (M. J. Gait, Ed.) pp 35-81, IRL Press, Washington, DC. (48) Manoharan, M., Tivel, K. L., and Cook, P. D. (1995) Lipidic Nucleic Acids. Tetrahedron Lett. 36, 3651-3654. (49) Forster, R. J. (1996) Reversibly Adsorbed Monolayers on Microelectrodes: Effect of Potential on the Adsorption Thermodynamics. Anal. Chem. 68, 3143. (50) Meunier-Prest, R., and Laviron, E. (1992) Electrochemical Reactions with Protonations at Equilibrium. Part XV. The 2-electron, 2-proton Bi-cubic Scheme. J. Electroanal. Chem. 328, 33-46. (51) Sletten, E., and Nerdal, W. (1997) Interaction of Mercury with Nucleic Acids and Their Components. In Metals in Biology (A. Sigel and H. Sigel, Eds.) Vol. 34, pp 479-501, Marcel Dekker, New York. (52) Ksenzhek, O. S., Petrova, S. A., Oleinik, S. V., Kolokyazhnyi, M. V., and Moskovskii, V. Z. (1977) Investigation of the redox properties of quinoid compounds. II. Anthraquinone and its derivatives. Soviet Electrochemistry 13, 151-158. (53) Salaita, G. N., and Hubbard, A. T. (1992) Adsorbed Organic Molecules at Well-Defined Electrode Surfaces. In Molecular Design of Electrode Surfaces, Techniques of Chemistry (R. W. Murray, Ed.) Vol. XXII, John Wiley & Sons, New York. (54) Bard, A. J., and Faulkner, L. R. (1980) In Electrochemical Methods: Fundamentals and Applications, John Wiley & Sons, New York. (55) Herne, T. M., and Tarlov, M. J. (1997) Characterization of DNA Probes Immobilized on Gold Surfaces. J. Am. Chem. Soc. 119, 8916-8920.

BC980095I